Abstract
The Transient Receptor Potential (TRP) family of ion channels is widely expressed in many cell types and plays various physiological roles. Growing evidence suggests that certain TRP channels are functionally expressed in the immune system. Indeed, an increasing number of reports have demonstrated the functional expression of several TRP channels in innate and adaptive immune cells and have highlighted their critical role in the activation and function of these cells. However, very few reviews have been entirely dedicated to this subject. Here, we will summarize the recent findings with regards to TRP channel expression in T cells and discuss their emerging role as regulators of T cell activation and functions. Moreover, these studies suggest that beyond their pharmaceutical interest in pain management, certain TRP channels may represent potential novel therapeutic targets for various immune-related diseases.
Keywords: T cells, T cell receptor, TRP channels, Ca2+ signaling
Introduction
This special issue of Seminars in Immunopathology focuses on “The role of TRP Ion Channels in Physiology and Pathology”. TRP channels have been studied for many years by neuroscientists, mostly as pain receptors [1]. Although the physiological roles of TRP channels are well characterized in the nervous system, growing evidence suggests that TRP channels are expressed and functional in a wide array of non-neuronal cells. Indeed, these channels are expressed in many organs and tissues, including brain, heart, liver, lung, small and large intestine, skeletal muscle, skin, pancreas and the immune system [2-4]. However, the role of TRP channels in immune cells including T cells is by far less understood.
The mammalian TRP family consists of 28 members divided into six subfamilies: 1 TRPA (Ankyrin), 7 TRPC (Canonical), 6 TRPV (Vanilloid), 8 TRPM (Melastatin), 3 TRPP (Polycystin) and 2 TRPML (Mucolipin) [5]. All TRP proteins have six putative transmembrane domains (S1-S6) with a cation-permeable pore region between S5 and S6 and functional TRP channels are formed by homo- or hetero-tetramers with N- and C-termini being located intracellularly [3, 5, 6]. Most TRP channels are non-selective cation channels with a permeability for Ca2+ over Na+ (ratio PCa/PNa) ranging from 0.1 to >100) [3, 6].
Ca2+ channels are mandatory molecules for T cell activation as they provide the Ca2+ influx necessary for the many signaling processes that direct cellular responses such as cytokine secretion, T cell proliferation and T cell differentiation in T helper (Th) effector cells [7-9]. A major target downstream of T cell receptor (TCR) activation is the phosphorylation of phospholipase Cγ1 (PLCγ1), which catalyzes the hydrolysis of phosphatidylinositol (4,5) bisphosphate (PIP2) into diacylglycerol (DAG) and inositol (1,4,5) trisphosphate (IP3). DAG and IP3 promote protein kinase C (PKC) activation and a release of Ca2+ from intracellular stores, respectively, followed by an influx of Ca2+ from the extracellular space, a process called store-operated Ca2+ entry (SOCE) [10]. Significant advances in our understanding of Ca2+ signaling in T cells and of the molecular machinery (i.e., ion channels) involved in this complex process have occurred over the last decade. One particular ion channel, the Ca2+ release-activated Ca2+ (CRAC) channel has been described as the major player in mediating SOCE in T cells [8, 11]. However, several additional families of ion channels are expressed at the plasma membrane of T cells and may therefore contribute to the regulation of Ca2+ entry in T cells [12-14]. Thus, before the discovery of ORAI1 as the pore-forming plasma membrane subunit of the CRAC channel [8, 11], several members of the TRP channel family (e.g., TRPC1/3, TRPV6) were considered as candidates to form the CRAC channel. However, the pore properties of most TRP channels studied appeared not to match those of the CRAC channel [3, 6]. Nevertheless, an increasing amount of literature has reported the contribution of several TRP channels to Ca2+ entry in T cells via different mechanisms: either directly via SOCE or non-SOCE pathways, or indirectly through the regulation of the driving force for Ca2+ via other Ca2+-permeable channels [12-14]. Although the evidences about TRP channel expression in T cells have been accumulating over the last decade, to the best of our knowledge only one review has been entirely dedicated to the expression and function of TRP channels in T cells so far [15].
In this review, we will attempt to summarize the current knowledge on the TRP channels that have being identified as functionally expressed in rodent or human T lymphocytes and discuss the mechanisms by which they regulate T cell activation and functions.
Evidences for TRP channel expression in T cells
The expression of a growing number of TRP channels has been reported to date in rodent and human T cells [15-17]. In most of these studies, analysis of TRP channel expression at the mRNA level using Northern blot, reverse transcriptase PCR (RT-PCR) or real-time quantitative PCR (q-PCR) techniques was frequently employed. However, analysis of protein expression and channel functionality (i.e., Ca2+ signaling and/or electrophysiological data) was not consistently performed. In addition, in many studies T cell lines but not primary T cells were used. Therefore, the functional expression of certain TRP family members in non-transformed cells remains to be determined. Finally, because of the relatively poor specificity of certain TRP agonists and antagonists, convincing studies rely on genetic approaches and the use of specific TRP knockout mouse or knockdown cells.
Despite the aforementioned limitations in the study of TRP channels in T cells, a certain number of studies have convincingly demonstrated the important role of TRP channels in T cell activation and functions. Among the TRP channels likely to be expressed and functional in T cells, to date the expression of TRPA1, TRPC1/2/3/5/6, TRPV1-6 and TRPM1/2/4/5/6/7 was reported by different investigators (Table 1).
Table 1. Expression of TRP channels reported to date in T cells.
| Channel | Cell type | Species | Detection methods | References | |
|---|---|---|---|---|---|
| TRPA1 | SP, LN, Jurkat | Human | NB, IB, IHC | [19] | |
| CD4+ | Mouse, Human | q-PCR, IB, IF, Ca2+, EP | [Bertin et al. Submitted] | ||
| TRPC1 | SP, LN, CD4+ | Mouse | RT-PCR | [16] | |
| CD4+, Jurkat | Human | RT-PCR, q-PCR | [17] | ||
| HPB-ALL, Jurkat | Human | RT-PCR, Ca2+ | [25] | ||
| TRPC2 | SP, LN, CD4+ | Mouse | RT-PCR | [16] | |
| TRPC3 | Jurkat | Human | SB, NB, IB, Ca2+, EP | [26] | |
| CD4+, Jurkat | Human | RT-PCR, q-PCR | [17] | ||
| Jurkat | Human | RT-PCR, q-PCR, Ca2+ | [30] | ||
| PBMC | Rat | IB, Ca2+ | [31] | ||
| TRPC5 | CD4+, CD8+ | Mouse | RT-PCR, IF, Ca2+ | [27, 28] | |
| TRPC6 | PBMC, Jurkat | Human | RT-PCR, IB, Ca2+ | [20] | |
| Jurkat | Human | RT-PCR, Ca2+ | [29] | ||
| Jurkat | Human | RT-PCR, q-PCR, Ca2+ | [30] | ||
| PBMC | Rat | IB, Ca2+ | [31] | ||
| TRPV1 | PBMC | Rat | RT-PCR | [36] | |
| CD3+ | Human | q-PCR | [37, 39] | ||
| PBMC | Human | RT-PCR, IF | [38] | ||
| CD4+, Jurkat | Human | RT-PCR, q-PCR | [17] | ||
| CD4+, Jurkat | Mouse, Human | q-PCR, IB, IF, FACS, Ca2+, EP | [40] | ||
| CD4+, Jurkat | Mouse, Human | IF, Ca2+ | [47] | ||
| Thymocytes | Rat | q-PCR, IB, IF, FACS, Ca2+ | [41, 42] | ||
| CD3+, Jurkat | Human | Ca2+, EP | [44, 45] | ||
| TRPV2 | PBMC | Human | RT-PCR, IF | [38] | |
| SP, LN, CD4+ | Mouse | RT-PCR | [16] | ||
| CD3+ | Human | q-PCR | [37, 39] | ||
| CD4+, Jurkat | Human | RT-PCR, q-PCR | [17] | ||
| CD4+, Jurkat | Mouse, Human | IF | [47] | ||
| CD4+, Jurkat | Human | q-PCR, Ca2+ | [48] | ||
| Jurkat | Human | EP | [49] | ||
| TRPV3 | SP, LN, CD4+ | Mouse | RT-PCR | [16] | |
| CD3+ | Human | q-PCR | [37, 39] | ||
| CD4+, Jurkat | Mouse, Human | IF | [47] | ||
| TRPV4 | SP, LN, CD4+ | Mouse | RT-PCR | [16] | |
| CD3+ | Human | q-PCR | [37, 39] | ||
| CD4+, Jurkat | Mouse, Human | IF, Ca2+ | [47] | ||
| TRPV5 | CD3+, Jurkat | Human | q-PCR, IB, EP | [50] | |
| Jurkat | Human | IF, Ca2+, EP | [57] | ||
| TRPV6 | Jurkat | Human | RT-PCR, q-PCR, Ca2+, EP | [51] | |
| CD3+, Jurkat | Human | q-PCR, IB, EP | [50] | ||
| Jurkat | Human | IF, Ca2+, EP | [57] | ||
| TRPM1 | SP, LN, CD4+ | Mouse | RT-PCR | [16] | |
| TRPM2 | CD4+, Jurkat | Human | RT-PCR, q-PCR | [17] | |
| SP, LN, CD4+ | Mouse | RT-PCR | [16] | ||
| Jurkat | Human | EP | [62] | ||
| Jurkat | Human | IB, IF, Ca2+, EP | [64] | ||
| CD3+, Jurkat | Human | IB, Ca2+, EP | [65] | ||
| CD4+ | Mouse | RT-PCR, IF | [66] | ||
| TRPM4 | SP, LN, CD4+ | Mouse | RT-PCR | [16] | |
| Jurkat | Human | RT-PCR, IB, Ca2+, EP | [67] | ||
| CD4+ | Mouse | IB, Ca2+ | [68] | ||
| TRPM5 | SP, LN, CD4+ | Mouse | RT-PCR | [16] | |
| TRPM6 | SP, LN, CD4+ | Mouse | RT-PCR | [16] | |
| TRPM7 | CD4+, Jurkat | Human | RT-PCR, q-PCR | [17] | |
| SP, LN, CD4+ | Mouse | RT-PCR | [16] | ||
| Thymocytes, CD3+ | Mouse | RT-PCR, IB, Mg2+, EP | [70] | ||
| CD3+, Jurkat | Mouse, Human | q-PCR, IB, EP | [72] | ||
| CD3+ | Human | IF, EP | [73] |
TRPA channels
The TRPA family contains only one member, TRPA1. This ion channel is best known in sensory neurons as a sensor for environmental irritants, inflammatory pain, and itching [18]. However, TRPA1 has a diverse tissue distribution and plays different roles in a variety of non-neuronal cells [2, 18]. TRPA1 expression at the mRNA and protein levels was reported in human secondary lymphoid organs (i.e., spleen and lymph nodes) as well as in a human T cell leukemic cell line (i.e., Jurkat) [19]. However, TRPA1 channel functionality was not investigated in this study. Recently, our group identified the constitutive expression of TRPA1 mRNA and protein in mouse and human primary CD4+ T cells. In addition, we demonstrated TRPA1 channel functionality at the plasma membrane of mouse CD4+ T cells by electrophysiological and Ca2+ imaging techniques. We identified that TRPA1 regulates TRPV1 channel activity in CD4+ T cells, and consequently controls CD4+ T cell activation and pro-inflammatory responses in models of colitis [Bertin et al. Submitted for publication, 2015].
TRPC channels
TRPC channels are subdivided into four subfamilies on the basis of sequence homology and similarities in function: TRPC1, TRPC2 (a pseudogene in human), TRPC3/6/7 and TRPC4/5. All TRPCs form non-selective ion channels with only a very modest Ca2+ selectivity [3, 6]. TRPC channels are widely expressed in the nervous system as well as in non-excitable cells [5]. Certain TRPC channels, specifically TRPC1, TRPC3, TRPC5 and TRPC6 are expressed in T cells [15, 20]. TRPC1 and TRPC3 have been proposed to act as store-operated channels and to interact with STIM1 and ORAI1 [21, 22]. Although this remains very controversial [8, 23, 24] and TRPC channels are no longer considered part of the CRAC channel, they may mediate other forms of Ca2+ influx relevant for T cell activation [17, 24]. For example, TRPC1 is functionally expressed in a human T cell line (HPB-ALL) and contributes to cannabinoid-induced Ca2+ influx in these cells independently of intracellular Ca2+ store depletion [25]. In another study, Jurkat cells with mutated TRPC3 channels showed decreased Ca2+ influx after TCR stimulation, which was rescue by overexpression of wild-type TRPC3 [26]. Furthermore, TRPC3 mRNA is strongly up-regulated in human primary CD4+ T cells following TCR stimulation and siRNA-mediated knockdown of TRPC3 led to a decrease in Ca2+ influx and proliferation [17].
Another subfamily member, TRPC5 is upregulated in activated mouse CD4+ and CD8+ effector T cells and mediates Ca2+ influx after binding of surface GM1 ganglioside by galectin-1 on regulatory T cells (Treg), potentially contributing to Treg-mediated suppression of effector T cell functions in experimental autoimmune encephalomyelitis (EAE) [27] and in the NOD (non-obese diabetic) mouse model of type-1 diabetes [28].
Tseng and colleagues reported that TRPC6 overexpression or antisense downregulation significantly alters the amplitude of phosphatidylinositol 3,4,5-trisphosphate (PIP3)- and TCR-induced Ca2+ responses in Jurkat cells [29]. In addition, silencing of TRPC3 and TRPC6 genes by shRNA abolished Ca2+ entry in Jurkat cells in response of certain free fatty acids [30]. Finally, a recent study utilizing a rat model of sepsis, demonstrated that TRPC3 and TRPC6 expression is upregulated in T cells and enhanced T cell apoptosis [31].
TRPV channels
Similarly to the TRPC family, the TRPV family can be divided into two subfamilies based on the structure, function and Ca2+ selectivity: TRPV1-4 are non-selective cation channels (PCa/PNa ≈ 1-10) whereas TRPV5/6 are exclusively permeable to Ca2+ (PCa/PNa > 100) [3, 6]. TRPVs are known to participate in thermosensation (e.g., TRPV1-4) [5], pain perception (e.g., TRPV1) [1] and Ca2+ entry in certain types of epithelial cells (e.g., TRPV1/5/6) [32, 33]. TRPV1 can associate with TRPV2 and TRPV3 and widespread interaction has been shown for TRPV1-TRPV4 [6, 34]. Furthermore, TRPV5 and TRPV6 can form heteromeric channel complexes [6, 35]. TRPV1 mRNA and protein expression was reported in primary mouse and human T cells by several different groups [17, 36-40]. TRPV1 mRNA and protein expression was also reported in mouse and rat thymocytes [41, 42]. The authors showed that capsaicin (a specific TRPV1 channel agonist) [43], increases intracellular Ca2+ concentration ([Ca2+]i) and induced thymocyte cell death via apoptotic or autophagic pathways in a dose-dependent manner [41, 42]. These results suggested that the TRPV1 channel could influence the development of thymocytes in rodents. A rise in [Ca2+]i and induction of apoptosis in response to capsaicin was also demonstrated in human peripheral T cells and in Jurkat cells [44]. However, although the Ca2+ mobilization appeared to be mediated by TRPV1, the induction of apoptosis was TRPV1-independent and was observed in response to prolonged and very high capsaicin concentration exposure (e.g., 200-250 mM) [44]. Functional TRPV1 expression in Jurkat cells was confirmed by another group using both Ca2+ imaging and whole-cell patch clamp [45]. Using capsazepine (a TRPV1 channel antagonist) [46], the authors showed that TRPV1 mediates the Ca2+ influx rise in response to phytohemagglutinin (PHA, a lectin that binds to glycosylated surface proteins including the TCR, and crosslinks them) and suggested that TRPV1 contributes to physiological TCR-induced Ca2+ influx and T cell activation. In line with these observations, our group recently demonstrated that TRPV1 is constitutively expressed in mouse and human primary CD4+ T cells where it acts as a non-store operated Ca2+ channel and contributes to TCR-induced Ca2+ influx [40]. Using Trpv1-/- mouse and mice with conditional overexpression of TRPV1 in CD4+ T cells, we identified that TRPV1 regulates downstream TCR-induced signaling, cytokine production and the pro-inflammatory properties of CD4+ T cells in models of T cell-mediated colitis. We also validated these findings in human primary CD4+ T cells using TRPV1-specific siRNA and channel antagonists. Very recently, Majhi et al., confirmed the critical role of TRPV1 in mediating TCR-induced Ca2+ influx and cytokine production in mouse and human primary T cells by using I-RTX (a specific TRPV1 antagonist) [46] and interestingly, also reported the protein expression of other TRPV subfamily members (i.e, TRPV2, TRPV3 and TRPV4) in these cells as well as in Jurkat [47].
TRPV2 mRNA and protein expression has been reported in mouse and human T cells previously [16, 17, 37-39]. Sauer and Jegla have reported the presence of TRPV2 in Jurkat cells and in primary human T cells in a patent application [48]. They identified that Jurkat cells treated with a dominant negative hTRPV2 or nucleofected with hTRPV2 shDNA display a defect in TCR- or thapsigargin-induced Ca2+ mobilization, predominantly during the sustained phase of Ca2+ influx. In addition, TRPV2 knockdown Jurkat cells displayed a reduction in Ca2+ store release similar to that induced by the knockdown of the lymphocyte-specific protein tyrosine kinase Lck [48]. More recently, Pottosin et al., published a detailed electrophysiological analysis of membrane stretch-activated currents in Jurkat cells and used TRPV2 siRNA to confirm the involvement of the TRPV2 channel in this process [49]. These studies therefore indicate that TRPV2 may play a role in Ca²+ signaling in lymphocytes and in its regulation by mechanical stress.
TRPV3 protein expression was recently reported in mouse and human primary T cells as well as in Jurkat [47]. However, in line with observations made by our group and by others, Trpv3 mRNA [16, 37, 40] and protein [47] are expressed at low level in T cells and Trpv3-/- mice don't have any obvious T cell phenotype [Bertin et al. Unpublished observations].
In contrast to TRPV3, TRPV4 appear to be present at a higher level in CD4+ T cells [16, 40, 47]. Interestingly, Majhi et al., reported a similar role for TRPV4 as for TRPV1 in contributing to TCR-induced Ca2+ influx and cytokine production in mouse and human T cells [47]. In accordance with the fact that TRPV1 and TRPV4 can form heteromeric channel complexes [34], these data suggest that inhibition of both TRPV1 and TRPV4 may have a synergistic effect in decreasing TCR-induced Ca2+ influx and consequently T cell activation.
The functional expression of the Ca2+-selective TRPV5 and TRPV6 channels was reported in mouse and human primary T cells and in Jurkat cells [15, 50, 51]. However, Trpv6-/- mice have no reported immunological phenotype [52] and TCR-induced Ca2+ influx in CD4+ T cells from Trpv6-/- mice is normal [53]. The TRPV6 channel, which is highly permeable to Ca2+, was shown to be triggered by store depletion [51] and, as certain TRPC subfamily members, was considered as a potential molecular entity of the CRAC channel. When a dominant-negative pore-region mutant of TRPV6 was expressed in Jurkat T cells, the CRAC current was found to be reduced [51]. However, the role of TRPV6 as a CRAC channel could not be established [54, 55] as BTP2 (a pyrazole derivative and CRAC channel inhibitor) did not affect the activity of the TRPV6 channel [56]. Recently, it was reported that both TRPV5 and TRPV6 are involved in Ca2+ entry and cell cycle progression of human primary T cells and Jurkat cells [50]. The activity of these two channels in Jurkat cells was also shown to be modulated by extracellular pH [57]. However, these conclusions rely on the use of Ruthenium red, a nonspecific channel inhibitor [58, 59]. Further, genetic evidences are still not comprehensive enough to fully demonstrate the physiological importance of TRPV5 and TRPV6 in the context of T cell activation and functions.
TRPM channels
The eight members of the mammalian TRPM subfamily are non-selective cation channels [5] and can be subdivided into two categories according to their Ca2+ permeability; practically impermeable (TRPM4/5) and moderately permeable (TRPM1/2/3/6/7/8) to Ca2+ [3, 6]. TRPMs are broadly expressed in both neuronal and non-neuronal cells and play diverse roles such as mediating direct Ca2+ influx, controlling Mg2+ entry and determining the membrane potential [5]. Among them, TRPM2, TRPM4 and TRPM7 appear to be constitutively expressed in CD4+ T cells [15, 17].
Sano et al., published one of the first reports on the role of TRPM2 (formerly LTRPC2) as a Ca2+ channel in T cells [60]. Their study demonstrated that intracellular secondary messengers ADP-ribose (ADPR) and nicotinamide adenine dinucleotide (NAD+) could directly stimulate TRPM2 channel activity to induce Ca2+ entry in Jurkat cells. Guse et al., previously demonstrated that a cyclic ADPR antagonist inhibits TCR-induced proliferation and expression of CD25 and HLA-DR in Jurkat cells [61]. Conversely, hydrogen peroxide (H2O2) can facilitate ADPR-mediated activation of TRPM2 and may therefore link TRPM2 activation to oxidative stress in various immune cells including T cells [62, 63]. Others showed that Concanavalin A (an antigen-independent mitogen) increases intracellular ADPR concentrations in Jurkat cells and that ADPR mediates Ca2+ influx through the TRPM2 channel and induces cell death [64]. In this study, the authors also showed that inhibition of ADPR formation or knockdown of TRPM2 both inhibited TRPM2-mediated Ca2+ influx. Another study similarly emphasized on the role of NAD+ and ADPR in mitogen-induced Ca2+ rise in Jurkat and in primary human T cells via TRPM2 [65]. More recently, Melzer and coworkers demonstrated that TRPM2 is expressed in mouse primary CD4+ T cells and contributes to T cell proliferation and production of pro-inflammatory cytokines after TCR stimulation [66]. The authors showed that the severity of EAE was reduced in Trpm2-/-mice and that this protection was associated with impaired TCR-stimulated proliferation as well as IFN-γ and IL-17 production by Trpm2-/- CD4+ T cells.
TRPM4 indirectly regulates Ca2+ entry by conducting Na+ and K+ currents [3, 6]. Interestingly, TRPM4 has been shown to negatively regulate SOCE in T cells by inducing membrane depolarization and by reducing the driving force for Ca2+ entry [67]. Launay et al., first detected TRPM4-mediated currents in Jurkat T cells and showed that downregulation of TRPM4 expression using an siRNA approach, increased PHA-induced Ca2+ entry and IL-2 production in Jurkat T cells [67]. More recently, Weber et al., demonstrated that TRPM4 is expressed at different levels in mouse Th cells (Th2 > Th1) and that it differentially regulates Ca2+ signaling and NFATc1 subcellular localization in these cells [68]. The authors found that inhibition of TRPM4 expression (via siRNA or a dominant negative construct) in Th2 cells resulted in increased Ca2+ influx, NFATc1 nuclear localization and IL-2 production as well as decreased cell motility and IL-4 production. Inhibition of TRPM4 expression in Th1 cells caused a decrease in Ca2+ influx and NFATc1 nuclear localization, resulting in an increase in cell motility, and a decrease in IL-2 and IFN-γ production [68]. These in vitro studies suggest an important role for TRPM4 in regulating T cell activation and differentiation in Th effector cells. However, in vivo validation of a role for TRPM4 in disease models has not yet been reported.
TRPM7 is a Mg2+-permeable, non-selective cation channel required for Mg2+ homeostasis in many cell types [69]. Since Trpm7-/- mice died prenatally, Clapham and colleagues used Lck-Cre mice to selectively delete TRPM7 in T cells [70]. The resulting mice with conditional deletion of TRPM7 in T cells displayed a severe defect in T cell development in the thymus. The authors noticed a block in transition from the double negative (CD4-CD8-) to double positive (CD4+CD8+) stage in Trpm7-/- thymocytes. As a result, both the number and the percentage of T cells in the periphery were reduced [70]. However, Trpm7-/- T cells did not show any significant defect in Mg2+ uptake suggesting that TRPM7 is dispensable for Mg2+ homeostasis in T cells and that another mechanism was responsible for the phenotype of these mice. Interestingly, TRPM7 also contains a regulatory serine-threonine kinase domain and is often referred to as a “chanzyme” for its channel-enzyme bifunctionality [71]. A recent study from the same group showed that Trpm7-/- T cells are deficient in Fas-mediated apoptosis and that this effect is dependent on TRPM7 activity as a channel rather than a kinase [72]. Therefore, the question arises whether the effect of TRPM7 deletion on T cell development is related to divalent cations other than Mg2+, such as Ca2+ or Na+. In line with this idea, TRPM7 was shown to regulate Ca2+ oscillations at the uropod of activated human primary T cells and to control their migration [73]. Finally, mice with conditional deletion of TRPM7 in T cells also displayed lung inflammation and pathophysiological similarity to mice in which the Fas receptor is selectively deleted in T cells, further supporting a role for TRP channels as T cell-intrinsic regulators of cell death [31, 41, 42, 44, 64] and inflammation [40, 74-76].
TRPP and TRPML channels
The TRPP and TRPML constitute two small subfamilies of non-selective cation channels that are predominantly expressed intracellularly [77]. TRPMLs play a critical role in endolysosomal trafficking and function [78] and genetic defects in TRPML1 and TRPML3 lead to a variety of cellular phenotypes including neurological and neurosensory deficiencies in humans and in animal models [77]. In addition, human TRPML2 has been shown to co-localize with major histocompatibility complex class I (MHC-I) and CD59 antigen [79], whereas mouse TRPML1 co-localizes with MHC-II in macrophage cell lines [80]. A role for TRPML1 and TRPML2 in vesicular transport in B cells has also been reported [81]. However, to the best of our knowledge their expression hasn't been reported to date in T cells and therefore remains to be investigated.
Modes of activation of TRP channels in T cells
TRP channels can respond to stimuli ranging from changes in temperature, pH, osmolarity, [Ca2+]i, to natural product compounds, pro-inflammatory agents, and various endogenous or exogenous stress mediators [5]. TRP channels are polymodal receptors and can therefore be activated via diverse mechanisms, some through direct binding of agonists (e.g., TRPV1 by capsaicin) [43], following receptor stimulation in a PLC- (e.g., TRPC1, TRPC4, TRPC5) or DAG-(e.g., TRPC3, TRPC6, TRPC7) dependent way [82, 83]. Temperature can also activate the thermosensitive TRP channels (e.g., TRPA1, TRPV1-4, TRPM8) [84] and store-depletion is discussed to be playing a role in modulating the activity of certain TRPC channels. Several studies have reported that STIM1 and ORAI1 can directly interact with TRPC channels, specifically with TRPC1 and TRPC3 [21, 22]. However as mentioned before, this remains a controversial discussion [8, 23, 24, 82].
The activity of certain TRP channels can also be modulated by cellular kinases (e.g., PKA, PKC, Src) [85, 86]. For example, our results suggested that phosphorylation by Lck is the gating mechanism of the TRPV1 channel in CD4+ T cells upon TCR stimulation [40, 87]. Tyrosine phosphorylation by Lck may activate TRPV1 downstream of TCR engagement by several potential mechanisms: (i) directly as it was shown for Src-mediated activation of TRPV1 [88] and other TRP family members such as TRPV4 [89], TRPV6 [90], TRPC3 [91, 92], TRPC6 [93] and TRPM2 [94]; (ii) indirectly by lowering it's activation threshold (i.e., sensitization) to endogenous agonists (e.g., DAG) [95]; or (iii) by inducing TRPV1 translocation from intracellular pools to the plasma membrane and increasing the surface density of TRPV1 channels [96-98].
Another mode of regulation of TRP channel activities resides in their multimerization capacity. Most TRPs can form heteromultimeric channel complexes with another subfamily member (e.g., TRPC1/5, TRPC4/5, TRPV5/6) [3, 5, 6, 35] or with a member of a different subfamily (e.g., TRPA1/V1) [99-101], creating a variety of chimeric channels with unique properties, as compared to homomultimers, where each TRP channel within a same heteromultimeric complex can influence each other's activity. For example, we recently identified that TRPA1 and TRPV1 are co-expressed in CD4+ T cells and that TRPA1 restrains TRPV1 channel activity in these cells. Indeed, Trpa1-/- CD4+ T cells display augmented TRPV1 channel activity and consequently, increased TCR-induced Ca2+ influx, T cell activation and pro-inflammatory capacities [Bertin et al. Submitted for publication, 2015].
In conclusion, TRP channels in T cells, such as in other cell types, appear to be regulated via several different mechanisms and to operate either as: (i) primary detectors (ligand-gated ion channel) of chemical and physical stimuli, (ii) as secondary transducers (downstream of a receptor e.g., TCR) of cell activation, or (iii) as ion transport channels.
Functional Roles of TRP Channels in T cells
Involvement in physiological T cell activation
The increase in [Ca2+]i is fundamental for T cell activation and biological functions [8, 9, 102, 103]. Like in other non-excitable cells, TRP channels in T cells contribute to Ca2+ influx either directly as Ca2+ permeable channels at the plasma membrane (e.g., TRPC1/3, TRPV1/4/5/6, TRPM2) or indirectly by controlling the membrane potential and thus, the driving force for Ca2+ entry through other Ca2+ entry channels (e.g., TMRP4) [3, 5, 12-14]. As mentioned above, several TRP channels appear to regulate the Ca2+ influx induced in response to TCR stimulation and therefore participate in the activation of T cells in physiological conditions. However, genetic evidence for the physiological roles of most TRP channels in T cells is still lacking in thoroughness. In comparison to CRAC, TRP channels may have mostly modulatory and homeostatic roles in T cells, which may be important when the immune system is challenged but not necessary apparent under normal homeostatic conditions in unchallenged mice. By mediating SOCE, CRAC is largely accepted as the major source of Ca2+ influx in T cells with certain TRP channels such as TRPC1, TRPC3 and TRPM4, acting as regulators of this process [8, 23, 24, 53]. For example, TRPM4 negatively regulates SOCE and plays an important physiological role in T cells by preventing Ca2+ overload [67]. Other TRP channels such as TRPV1, TRPV2 and TRPV4 may act independently of CRAC as non-store operated channels in T cells [40, 47, 48] but nevertheless appear to contribute to the physiological Ca2+ influx following TCR stimulation. Despite the significant advances in the understanding of Ca2+ signaling in T cells, further studies are needed to determine the relative contribution of TRP channels to the overall Ca2+ entry process that follows engagement of the TCR [103].
Certain TRP channels may also play an important physiological role in T cells by modulating T cell activation in response of temperature changes. Although the effect of temperature on the immune response is well documented [104-107], the exact molecular mechanisms are poorly understood [108]. The expression of thermosensitive TRPs such as TRPA1, TRPV1-4 and TRPM2 in T cells suggest that these channels may play a role in the effect of temperature changes on T cell activation and ensuing immune responses.
Finally, by regulating Ca2+ signals in T cells, TRP channels may be important regulator of cell survival and cell death. Although an entry of Ca2+ is vital for cell functions and so are necessary for cell survival, an excess of [Ca2+]i can have detrimental effects. Recently it has become clear that cellular Ca2+ overload, or perturbation of intracellular Ca2+ compartmentalization, can cause cytotoxicity and trigger cell death via apoptotic, necrotic or autophagic pathways [109]. As documented for TRPV1 [41, 42] and TRPM2 [64] in response to chemical or oxidative stress respectively, and for TRPM4 in response to SOCE [67], TRP channels may play a critical role in the process of Ca2+ overload in T cells under physiological and pathological conditions.
Involvement in pathological T cell activation
Defects in ion channel function have been linked to various diseases, so-called channelopathies. However, among the 28 mammalian TRP channels, relatively few have been directly connected to human diseases [110]. As mentioned above, it is now clear that several TRP channels are functionally expressed in T cells and their activation contributes to T cell activation including proliferation, differentiation into effector cells and release of specific inflammatory cytokines, suggesting a potential role for these ion channels in immune-related diseases [4, 111]. Changes in TRP channel expression or function in T cells may be associated with a variety of autoimmune and chronic inflammatory diseases, such as asthma [112-114], inflammatory bowel disease [40, 76, 115, 116], rheumatoid arthritis [117, 118], multiple sclerosis [27, 66, 119], type-1 diabetes [4, 28], lupus [120], and psoriasis [4] but in vivo studies focusing on the role of TRP channels in T cells are limited and up to now restricted to TRPV1 [40], TRPC5 [27], TRPC6 [112], TRPM2 [66] and TRPM7 [70, 72]. The above-mentioned studies therefore suggest that certain TRP channels could represent new drug targets for the management of various T cell-mediated diseases. In addition, such as other molecules interfering with Ca2+ signaling in T cells (e.g., cyclosporin A and FK506, two calcineurin inhibitors), certain TRP channels modulators may have potential therapeutic applications in organ transplantation, where T cells are key players in the process of graft rejection and transplantation tolerance [121].
Conclusions
It is becoming clear that T cell functions are regulated by a network of different ion channels including CRAC, TRPs, voltage-gated Ca2+ (Cav) channels, P2X receptors, Ca2+-activated K+ channels (KCa) and voltage-gated K+ (Kv) channels [12-14, 102]. However, compared to the CRAC channel, the contribution of other families of ion channels to TCR-induced Ca2+ influx and T cell functions has not been investigated as comprehensively.
In this review, we attempted to summarize the recent studies that demonstrate the functional expression and the critical role of TRP channels in T cells. Despite the increasing number of studies reporting the expression of various TRP channels at the mRNA and/or protein level in T cells, only a few have demonstrated the functionality of TRP channels in primary T cells. In addition, reports using conditional mice with T cell-specific deletion of Trp genes are restricted until now to TRPM7 [70, 72] and most studies have employed T cells isolated from mice with ubiquitous inactivation of individual Trp genes in which the observed phenotype may potentially be affected by developmental defects or compensatory upregulation of other genes in adult animals. Therefore, more in vivo studies with conditional TRP-deficient mice are needed in addition to the use of si/shRNA-mediated knockdown strategies in in vitro experiments with primary T cells in order to unambiguously demonstrate the cell-intrinsic role of TRP channels in T cells. In spite of these limitations, the most important conclusion of this review is that several TRP channels are functionally expressed in T cells and contribute to T cell activation under physiological and pathological conditions. However, how TRP channels function in T cells and how they interact with other family members and with other channels (e.g., CRAC channel) remain poorly understood. Future studies will be needed to explore the complex interplay between ion channels in T cells and to identify the precise role of each channel during T cell development and in the different effectors T cell subsets.
Acknowledgments
We apologize to the colleagues whose work could not be cited due to space limitations or may have been omitted. We thank Hannah Federman for proofreading the manuscript. This work was supported by a grant from the NIH (U01 AI095623), an award to E.R. from the Crohn's and Colitis Foundation of America (CCFA) (SRA#330251), and a research fellowship to S.B. from the CCFA (RFA#3574).
Abbreviations
- Ca2+
Ca2+ imaging
- CD3+
primary CD3+ T cells
- CD4+
primary CD4+ T cells
- CD8+
primary CD8+ T cells
- EP
electrophysiology
- IB
immunoblot
- IHC
immunohistochemistry
- LN
lymph nodes
- Mg2+
Mg2+ imaging
- NB
northern blot
- PBMC
peripheral blood mononuclear cells
- q-PCR
quantitative PCR
- RT-PCR
conventional reverse transcription PCR
- SB
southern blot
- SP
spleen
Footnotes
Author's contribution S.B. and E.R. wrote this review.
Competing Financial Interest The authors declare no competing financial interests.
References
- 1.Julius D. TRP channels and pain. Annu Rev Cell Dev Biol. 2013;29:355–84. doi: 10.1146/annurev-cellbio-101011-155833. [DOI] [PubMed] [Google Scholar]
- 2.Fernandes ES, Fernandes MA, Keeble JE. The functions of TRPA1 and TRPV1: moving away from sensory nerves. Br J Pharmacol. 2012;166:510–21. doi: 10.1111/j.1476-5381.2012.01851.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Gees M, Colsoul B, Nilius B. The role of transient receptor potential cation channels in Ca2+ signaling. Cold Spring Harb Perspect Biol. 2010;2:a003962. doi: 10.1101/cshperspect.a003962. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Nilius B, Szallasi A. Transient receptor potential channels as drug targets: from the science of basic research to the art of medicine. Pharmacol Rev. 2014;66:676–814. doi: 10.1124/pr.113.008268. [DOI] [PubMed] [Google Scholar]
- 5.Venkatachalam K, Montell C. TRP channels. Annu Rev Biochem. 2007;76:387–41. doi: 10.1146/annurev.biochem.75.103004.142819. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Owsianik G, Talavera K, Voets T, Nilius B. Permeation and selectivity of TRP channels. Annu Rev Physiol. 2006;68:685–717. doi: 10.1146/annurev.physiol.68.040204.101406. [DOI] [PubMed] [Google Scholar]
- 7.Dolmetsch RE, Lewis RS, Goodnow CC, Healy JI. Differential activation of transcription factors induced by Ca2+ response amplitude and duration. Nature. 1997;386:855–8. doi: 10.1038/386855a0. [DOI] [PubMed] [Google Scholar]
- 8.Hogan PG, Lewis RS, Rao A. Molecular basis of calcium signaling in lymphocytes: STIM and ORAI. Annu Rev Immunol. 2010;28:491–533. doi: 10.1146/annurev.immunol.021908.132550. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Vig M, Kinet JP. Calcium signaling in immune cells. Nat Immunol. 2009;10:21–7. doi: 10.1038/ni.f.220. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Parekh AB, Penner R. Store depletion and calcium influx. Physiol Rev. 1997;77:901–30. doi: 10.1152/physrev.1997.77.4.901. [DOI] [PubMed] [Google Scholar]
- 11.Oh-hora M, Rao A. Calcium signaling in lymphocytes. Curr Opin Immunol. 2008;20:250–8. doi: 10.1016/j.coi.2008.04.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Feske S, Skolnik EY, Prakriya M. Ion channels and transporters in lymphocyte function and immunity. Nat Rev Immunol. 2012;12:532–47. doi: 10.1038/nri3233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Badou A, Jha MK, Matza D, Flavell RA. Emerging Roles of L-Type Voltage-Gated and Other Calcium Channels in T Lymphocytes. Front Immunol. 2013;4:243. doi: 10.3389/fimmu.2013.00243. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Nohara LL, Stanwood SR, Omilusik KD, Jefferies WA. Tweeters, Woofers and Horns: The Complex Orchestration of Calcium Currents in T Lymphocytes. Front Immunol. 2015;6:234. doi: 10.3389/fimmu.2015.00234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Schwarz EC, Wolfs MJ, Tonner S, Wenning AS, Quintana A, Griesemer D, Hoth M. TRP channels in lymphocytes. Handb Exp Pharmacol. 2007;179:445–56. doi: 10.1007/978-3-540-34891-7_26. [DOI] [PubMed] [Google Scholar]
- 16.Inada H, Iida T, Tominaga M. Different expression patterns of TRP genes in murine B and T lymphocytes. Biochem Biophys Res Commun. 2006;350:762–7. doi: 10.1016/j.bbrc.2006.09.111. [DOI] [PubMed] [Google Scholar]
- 17.Wenning AS, Neblung K, Strauss B, Wolfs MJ, Sappok A, Hoth M, Schwarz EC. TRP expression pattern and the functional importance of TRPC3 in primary human T-cells. Biochim Biophys Acta. 2011;1813:412–23. doi: 10.1016/j.bbamcr.2010.12.022. [DOI] [PubMed] [Google Scholar]
- 18.Nilius B, Appendino G, Owsianik G. The transient receptor potential channel TRPA1: from gene to pathophysiology. Pflugers Arch. 2012;464:425–58. doi: 10.1007/s00424-012-1158-z. [DOI] [PubMed] [Google Scholar]
- 19.Stokes A, Wakano C, Koblan-Huberson M, Adra CN, Fleig A, Turner H. TRPA1 is a substrate for de-ubiquitination by the tumor suppressor CYLD. Cell Signal. 2006;18:1584–9. doi: 10.1016/j.cellsig.2005.12.009. [DOI] [PubMed] [Google Scholar]
- 20.Gamberucci A, Giurisato E, Pizzo P, Tassi M, Giunti R, McIntosh DP, Benedetti A. Diacylglycerol activates the influx of extracellular cations in T-lymphocytes independently of intracellular calcium-store depletion and possibly involving endogenous TRP6 gene products. Biochem J. 2002;364:245–54. doi: 10.1042/bj3640245. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Potier M, Trebak M. New developments in the signaling mechanisms of the store-operated calcium entry pathway. Pflugers Arch. 2008;457:405–1. doi: 10.1007/s00424-008-0533-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Cheng KT, Ong HL, Liu X, Ambudkar IS. Contribution and regulation of TRPC channels in store-operated Ca2+ entry. Curr Top Membr. 2013;71:149–7. doi: 10.1016/B978-0-12-407870-3.00007-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Kim MS, Zeng W, Yuan JP, Shin DM, Worley PF, Muallem S. Native Store-operated Ca2+ Influx Requires the Channel Function of Orai1 and TRPC1. J Biol Chem. 2009;284:9733–41. doi: 10.1074/jbc.M808097200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Saul S, Stanisz H, Backes CS, Schwarz EC, Hoth M. How ORAI and TRP channels interfere with each other: interaction models and examples from the immune system and the skin. Eur J Pharmacol. 2014;739:49–5. doi: 10.1016/j.ejphar.2013.10.071. [DOI] [PubMed] [Google Scholar]
- 25.Rao GK, Kaminski NE. Induction of intracellular calcium elevation by Delta9-tetrahydrocannabinol in T cells involves TRPC1 channels. J Leukoc Biol. 2006;79:202–13. doi: 10.1189/jlb.0505274. [DOI] [PubMed] [Google Scholar]
- 26.Philipp S, Strauss B, Hirnet D, Wissenbach U, Mery L, Flockerzi V, Hoth M. TRPC3 mediates T-cell receptor-dependent calcium entry in human T-lymphocytes. J Biol Chem. 2003;278:26629–38. doi: 10.1074/jbc.M304044200. [DOI] [PubMed] [Google Scholar]
- 27.Wang J, Lu ZH, Gabius HJ, Rohowsky-Kochan C, Ledeen RW, Wu G. Cross-linking of GM1 ganglioside by galectin-1 mediates regulatory T cell activity involving TRPC5 channel activation: possible role in suppressing experimental autoimmune encephalomyelitis. J Immunol. 2009;182:4036–45. doi: 10.4049/jimmunol.0802981. [DOI] [PubMed] [Google Scholar]
- 28.Wu G, Lu ZH, Gabius HJ, Ledeen RW, Bleich D. Ganglioside GM1 deficiency in effector T cells from NOD mice induces resistance to regulatory T-cell suppression. Diabetes. 2011;60:2341–9. doi: 10.2337/db10-1309. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Tseng PH, Lin HP, Hu H, Wang C, Zhu MX, Chen CS. The canonical transient receptor potential 6 channel as a putative phosphatidylinositol 3,4,5-trisphosphate-sensitive calcium entry system. Biochemistry. 2004;43:11701–8. doi: 10.1021/bi049349f. [DOI] [PubMed] [Google Scholar]
- 30.Carrillo C, Hichami A, Andreoletti P, Cherkaoui-Malki M, del Mar Cavia M, Abdoul-Azize S, Alonso-Torre SR, Khan NA. Diacylglycerol-containing oleic acid induces increases in [Ca(2+)] (i) via TRPC3/6 channels in human T-cells. Biochim Biophys Acta. 2012;1821:618–26. doi: 10.1016/j.bbalip.2012.01.008. [DOI] [PubMed] [Google Scholar]
- 31.Wu QY, Sun MR, Wu CL, Li Y, Du JJ, Zeng JY, Bi HL, Sun YH. Activation of calcium-sensing receptor increases TRPC3/6 expression in T lymphocyte in sepsis. Mol Immunol. 2015;64:18–25. doi: 10.1016/j.molimm.2014.10.018. [DOI] [PubMed] [Google Scholar]
- 32.de Jong PR, Takahashi N, Harris AR, Lee J, Bertin S, Jeffries J, Jung M, Duong J, Triano AI, Lee J, Niv Y, Herdman DS, Taniguchi K, Kim CW, Dong H, Eckmann L, Stanford SM, Bottini N, Corr M, Raz E. Ion channel TRPV1-dependent activation of PTP1B suppresses EGFR-associated intestinal tumorigenesis. J Clin Invest. 2014;124:3793–806. doi: 10.1172/JCI72340. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.den Dekker E, Hoenderop JG, Nilius B, Bindels RJ. The epithelial calcium channels, TRPV5 & TRPV6: from identification towards regulation. Cell Calcium. 2003;33:497–50. doi: 10.1016/s0143-4160(03)00065-4. [DOI] [PubMed] [Google Scholar]
- 34.Hellwig N, Albrecht N, Harteneck C, Schultz G, Schaefer M. Homo- and heteromeric assembly of TRPV channel subunits. J Cell Sci. 2005;118:917–28. doi: 10.1242/jcs.01675. [DOI] [PubMed] [Google Scholar]
- 35.Hoenderop JG, Voets T, Hoefs S, Weidema F, Prenen J, Nilius B, Bindels RJ. Homo- and heterotetrameric architecture of the epithelial Ca2+ channels TRPV5 and TRPV6. EMBO J. 2003;22:776–85. doi: 10.1093/emboj/cdg080. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Schumacher MA, Moff I, Sudanagunta SP, Levine JD. Molecular cloning of an N-terminal splice variant of the capsaicin receptor. Loss of N-terminal domain suggests functional divergence among capsaicin receptor subtypes. J Biol Chem. 2000;275:2756–62. doi: 10.1074/jbc.275.4.2756. [DOI] [PubMed] [Google Scholar]
- 37.Spinsanti G, Zannolli R, Panti C, Ceccarelli I, Marsili L, Bachiocco V, Frati F, Aloisi AM. Quantitative Real-Time PCR detection of TRPV1-4 gene expression in human leukocytes from healthy and hyposensitive subjects. Mol Pain. 2008;4:51. doi: 10.1186/1744-8069-4-51. 8069-4-51. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Saunders CI, Kunde DA, Crawford A, Geraghty DP. Expression of transient receptor potential vanilloid 1 (TRPV1) and 2 (TRPV2) in human peripheral blood. Mol Immunol. 2007;44:1429–35. doi: 10.1016/j.molimm.2006.04.027. [DOI] [PubMed] [Google Scholar]
- 39.Bachiocco V, Bergamaschi R, Spinsanti G, Lima M, Romagnoli R, Sorda G, Aloisi AM. Lymphocyte TRPV 1-4 gene expression and MIF blood levels in a young girl clinically diagnosed with HSAN IV. Clin J Pain. 2011;27:631–4. doi: 10.1097/AJP.0b013e3182119356. [DOI] [PubMed] [Google Scholar]
- 40.Bertin S, Aoki-Nonaka Y, de Jong PR, Nohara LL, Xu H, Stanwood SR, Srikanth S, Lee J, To K, Abramson L, Yu T, Han T, Touma R, Li X, Gonzalez-Navajas JM, Herdman S, Corr M, Fu G, Dong H, Gwack Y, Franco A, Jefferies WA, Raz E. The ion channel TRPV1 regulates the activation and proinflammatory properties of CD4(+) T cells. Nat Immunol. 2014;15:1055–63. doi: 10.1038/ni.3009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Amantini C, Mosca M, Lucciarini R, Perfumi M, Morrone S, Piccoli M, Santoni G. Distinct thymocyte subsets express the vanilloid receptor VR1 that mediates capsaicin-induced apoptotic cell death. Cell Death Differ. 2004;11:1342–56. doi: 10.1038/sj.cdd.4401506. [DOI] [PubMed] [Google Scholar]
- 42.Farfariello V, Amantini C, Santoni G. Transient receptor potential vanilloid 1 activation induces autophagy in thymocytes through ROS-regulated AMPK and Atg4C pathways. J Leukoc Biol. 2012;92:421–3. doi: 10.1189/jlb.0312123. [DOI] [PubMed] [Google Scholar]
- 43.Caterina MJ, Schumacher MA, Tominaga M, Rosen TA, Levine JD, Julius D. The capsaicin receptor: a heat-activated ion channel in the pain pathway. Nature. 1997;389:816–24. doi: 10.1038/39807. [DOI] [PubMed] [Google Scholar]
- 44.Macho A, Calzado MA, Munoz-Blanco J, Gomez-Diaz C, Gajate C, Mollinedo F, Navas P, Munoz E. Selective induction of apoptosis by capsaicin in transformed cells: the role of reactive oxygen species and calcium. Cell Death Differ. 1999;6:155–6. doi: 10.1038/sj.cdd.4400465. [DOI] [PubMed] [Google Scholar]
- 45.Yu Y, Carter CR, Youssef N, Dyck JR, Light PE. Intracellular long-chain acyl CoAs activate TRPV1 channels. PLoS One. 2014;9:e96597. doi: 10.1371/journal.pone.0096597. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Szallasi A, Cortright DN, Blum CA, Eid SR. The vanilloid receptor TRPV1: 10 years from channel cloning to antagonist proof-of-concept. Nat Rev Drug Discov. 2007;6:357–72. doi: 10.1038/nrd2280. [DOI] [PubMed] [Google Scholar]
- 47.Majhi RK, Sahoo SS, Yadav M, Pratheek BM, Chattopadhyay S, Goswami C. Functional expression of TRPV channels in T cells and their implications in immune regulation. FEBS J. 2015;282:2661–8. doi: 10.1111/febs.13306. [DOI] [PubMed] [Google Scholar]
- 48.James JT, Karsten S. Methods for identifying t cell activation-modulating compounds 2006 [Google Scholar]
- 49.Pottosin I, Delgado-Enciso I, Bonales-Alatorre E, Nieto-Pescador MG, Moreno-Galindo EG, Dobrovinskaya O. Mechanosensitive Ca(2)(+)-permeable channels in human leukemic cells: pharmacological and molecular evidence for TRPV2. Biochim Biophys Acta. 2015;1848:51–9. doi: 10.1016/j.bbamem.2014.09.008. [DOI] [PubMed] [Google Scholar]
- 50.Vassilieva IO, Tomilin VN, Marakhova II, Shatrova AN, Negulyaev YA, Semenova SB. Expression of transient receptor potential vanilloid channels TRPV5 and TRPV6 in human blood lymphocytes and Jurkat leukemia T cells. J Membr Biol. 2013;246:131–40. doi: 10.1007/s00232-012-9511-x. [DOI] [PubMed] [Google Scholar]
- 51.Cui J, Bian JS, Kagan A, McDonald TV. CaT1 contributes to the stores-operated calcium current in Jurkat T-lymphocytes. J Biol Chem. 2002;277:47175–83. doi: 10.1074/jbc.M205870200. [DOI] [PubMed] [Google Scholar]
- 52.Suzuki Y, Kovacs CS, Takanaga H, Peng JB, Landowski CP, Hediger MA. Calcium channel TRPV6 is involved in murine maternal-fetal calcium transport. J Bone Miner Res. 2008;23:1249–56. doi: 10.1359/JBMR.080314. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Feske S, Wulff H, Skolnik EY. Ion channels in innate and adaptive immunity. Annu Rev Immunol. 2015;33:291–353. doi: 10.1146/annurev-immunol-032414-112212. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Voets T, Prenen J, Fleig A, Vennekens R, Watanabe H, Hoenderop JG, Bindels RJ, Droogmans G, Penner R, Nilius B. CaT1 and the calcium release-activated calcium channel manifest distinct pore properties. J Biol Chem. 2001;276:47767–70. doi: 10.1074/jbc.C100607200. [DOI] [PubMed] [Google Scholar]
- 55.Bodding M, Wissenbach U, Flockerzi V. The recombinant human TRPV6 channel functions as Ca2+ sensor in human embryonic kidney and rat basophilic leukemia cells. J Biol Chem. 2002;277:36656–64. doi: 10.1074/jbc.M202822200. [DOI] [PubMed] [Google Scholar]
- 56.Schwarz EC, Wissenbach U, Niemeyer BA, Strauss B, Philipp SE, Flockerzi V, Hoth M. TRPV6 potentiates calcium-dependent cell proliferation. Cell Calcium. 2006;39:163–7. doi: 10.1016/j.ceca.2005.10.006. [DOI] [PubMed] [Google Scholar]
- 57.Tomilin VN, Cherezova AL, Negulyaev YA, Semenova SB. TRPV5/V6 Channels Mediate Ca Influx in Jurkat T Cells Under the Control of Extracellular pH. J Cell Biochem. 2015 doi: 10.1002/jcb.25264. [DOI] [PubMed] [Google Scholar]
- 58.Charuk JH, Pirraglia CA, Reithmeier RA. Interaction of ruthenium red with Ca2(+)-binding proteins. Anal Biochem. 1990;188:123–31. doi: 10.1016/0003-2697(90)90539-l. [DOI] [PubMed] [Google Scholar]
- 59.Vriens J, Appendino G, Nilius B. Pharmacology of vanilloid transient receptor potential cation channels. Mol Pharmacol. 2009;75:1262–79. doi: 10.1124/mol.109.055624. [DOI] [PubMed] [Google Scholar]
- 60.Sano Y, Inamura K, Miyake A, Mochizuki S, Yokoi H, Matsushime H, Furuichi K. Immunocyte Ca2+ influx system mediated by LTRPC2. Science. 2001;293:1327–30. doi: 10.1126/science.1062473. [DOI] [PubMed] [Google Scholar]
- 61.Guse AH, da Silva CP, Berg I, Skapenko AL, Weber K, Heyer P, Hohenegger M, Ashamu GA, Schulze-Koops H, Potter BV, Mayr GW. Regulation of calcium signalling in T lymphocytes by the second messenger cyclic ADP-ribose. Nature. 1999;398:70–3. doi: 10.1038/18024. [DOI] [PubMed] [Google Scholar]
- 62.Beck A, Kolisek M, Bagley LA, Fleig A, Penner R. Nicotinic acid adenine dinucleotide phosphate and cyclic ADP-ribose regulate TRPM2 channels in T lymphocytes. FASEB J. 2006;20:962–4. doi: 10.1096/fj.05-5538fje. [DOI] [PubMed] [Google Scholar]
- 63.Syed Mortadza SA, Wang L, Li D, Jiang LH. TRPM2 Channel-Mediated ROS-Sensitive Ca(2+) Signaling Mechanisms in Immune Cells. Front Immunol. 2015;6:407. doi: 10.3389/fimmu.2015.00407. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Gasser A, Glassmeier G, Fliegert R, Langhorst MF, Meinke S, Hein D, Kruger S, Weber K, Heiner I, Oppenheimer N, Schwarz JR, Guse AH. Activation of T cell calcium influx by the second messenger ADP-ribose. J Biol Chem. 2006;281:2489–96. doi: 10.1074/jbc.M506525200. [DOI] [PubMed] [Google Scholar]
- 65.Magnone M, Bauer I, Poggi A, Mannino E, Sturla L, Brini M, Zocchi E, De Flora A, Nencioni A, Bruzzone S. NAD+ levels control Ca2+ store replenishment and mitogen-induced increase of cytosolic Ca2+ by Cyclic ADP-ribose-dependent TRPM2 channel gating in human T lymphocytes. J Biol Chem. 2012;287:21067–81. doi: 10.1074/jbc.M111.324269. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Melzer N, Hicking G, Gobel K, Wiendl H. TRPM2 cation channels modulate T cell effector functions and contribute to autoimmune CNS inflammation. PLoS One. 2012;7:e47617. doi: 10.1371/journal.pone.0047617. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Launay P, Cheng H, Srivatsan S, Penner R, Fleig A, Kinet JP. TRPM4 regulates calcium oscillations after T cell activation. Science. 2004;306:1374–7. doi: 10.1126/science.1098845. [DOI] [PubMed] [Google Scholar]
- 68.Weber KS, Hildner K, Murphy KM, Allen PM. Trpm4 differentially regulates Th1 and Th2 function by altering calcium signaling and NFAT localization. J Immunol. 2010;185:2836–4. doi: 10.4049/jimmunol.1000880. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Kerschbaum HH, Kozak JA, Cahalan MD. Polyvalent cations as permeant probes of MIC and TRPM7 pores. Biophys J. 2003;84:2293–305. doi: 10.1016/S0006-3495(03)75035-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Jin J, Desai BN, Navarro B, Donovan A, Andrews NC, Clapham DE. Deletion of Trpm7 disrupts embryonic development and thymopoiesis without altering Mg2+ homeostasis. Science. 2008;322:756–60. doi: 10.1126/science.1163493. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Runnels LW, Yue L, Clapham DE. TRP-PLIK, a bifunctional protein with kinase and ion channel activities. Science. 2001;291:1043–7. doi: 10.1126/science.1058519. [DOI] [PubMed] [Google Scholar]
- 72.Desai BN, Krapivinsky G, Navarro B, Krapivinsky L, Carter BC, Febvay S, Delling M, Penumaka A, Ramsey IS, Manasian Y, Clapham DE. Cleavage of TRPM7 releases the kinase domain from the ion channel and regulates its participation in Fas-induced apoptosis. Dev Cell. 2012;22:1149–62. doi: 10.1016/j.devcel.2012.04.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Kuras Z, Yun YH, Chimote AA, Neumeier L, Conforti L. KCa3.1 and TRPM7 channels at the uropod regulate migration of activated human T cells. PLoS One. 2012;7:e43859. doi: 10.1371/journal.pone.0043859. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Barbet G, Demion M, Moura IC, Serafini N, Leger T, Vrtovsnik F, Monteiro RC, Guinamard R, Kinet JP, Launay P. The calcium-activated nonselective cation channel TRPM4 is essential for the migration but not the maturation of dendritic cells. Nat Immunol. 2008;9:1148–56. doi: 10.1038/ni.1648. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Link TM, Park U, Vonakis BM, Raben DM, Soloski MJ, Caterina MJ. TRPV2 has a pivotal role in macrophage particle binding and phagocytosis. Nat Immunol. 2010;11:232–9. doi: 10.1038/ni.1842. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Yamamoto S, Shimizu S, Kiyonaka S, Takahashi N, Wajima T, Hara Y, Negoro T, Hiroi T, Kiuchi Y, Okada T, Kaneko S, Lange I, Fleig A, Penner R, Nishi M, Takeshima H, Mori Y. TRPM2-mediated Ca2+influx induces chemokine production in monocytes that aggravates inflammatory neutrophil infiltration. Nat Med. 2008;14:738–47. doi: 10.1038/nm1758. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Qian F, Noben-Trauth K. Cellular and molecular function of mucolipins (TRPML) and polycystin 2 (TRPP2) Pflugers Arch. 2005;451:277–85. doi: 10.1007/s00424-005-1469-4. [DOI] [PubMed] [Google Scholar]
- 78.Venkatachalam K, Wong CO, Zhu MX. The role of TRPMLs in endolysosomal trafficking and function. Cell Calcium. 2015;58:48–56. doi: 10.1016/j.ceca.2014.10.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Karacsonyi C, Miguel AS, Puertollano R. Mucolipin-2 localizes to the Arf6-associated pathway and regulates recycling of GPI-APs. Traffic. 2007;8:1404–1. doi: 10.1111/j.1600-0854.2007.00619.x. [DOI] [PubMed] [Google Scholar]
- 80.Thompson EG, Schaheen L, Dang H, Fares H. Lysosomal trafficking functions of mucolipin-1 in murine macrophages. BMC Cell Biol. 2007;8:54. doi: 10.1186/1471-2121-8-54. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Song Y, Dayalu R, Matthews SA, Scharenberg AM. TRPML cation channels regulate the specialized lysosomal compartment of vertebrate B-lymphocytes. Eur J Cell Biol. 2006;85:1253–64. doi: 10.1016/j.ejcb.2006.08.004. [DOI] [PubMed] [Google Scholar]
- 82.Dietrich A, Kalwa H, Gudermann T. TRPC channels in vascular cell function. Thromb Haemost. 2010;103:262–70. doi: 10.1160/TH09-08-0517. [DOI] [PubMed] [Google Scholar]
- 83.Rohacs T, Nilius B. Regulation of transient receptor potential (TRP) channels by phosphoinositides. Pflugers Arch. 2007;455:157–68. doi: 10.1007/s00424-007-0275-6. [DOI] [PubMed] [Google Scholar]
- 84.Hilton JK, Rath P, Helsell CV, Beckstein O, Van Horn WD. Understanding thermosensitive transient receptor potential channels as versatile polymodal cellular sensors. Biochemistry. 2015;54:2401–13. doi: 10.1021/acs.biochem.5b00071. [DOI] [PubMed] [Google Scholar]
- 85.Yao X, Kwan HY, Huang Y. Regulation of TRP channels by phosphorylation. Neurosignals. 2005;14:273–80. doi: 10.1159/000093042. [DOI] [PubMed] [Google Scholar]
- 86.Voolstra O, Huber A. Post-Translational Modifications of TRP Channels. Cells. 2014;3:258–87. doi: 10.3390/cells3020258. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Bertin S, de Jong PR, Jefferies WA, Raz E. Novel immune function for the TRPV1 channel in T lymphocytes. Channels (Austin) 2014;8:479–80. doi: 10.4161/19336950.2014.991640. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Jin X, Morsy N, Winston J, Pasricha PJ, Garrett K, Akbarali HI. Modulation of TRPV1 by nonreceptor tyrosine kinase, c-Src kinase. Am J Physiol Cell Physiol. 2004;287:C558–63. doi: 10.1152/ajpcell.00113.2004. [DOI] [PubMed] [Google Scholar]
- 89.Xu H, Zhao H, Tian W, Yoshida K, Roullet JB, Cohen DM. Regulation of a transient receptor potential (TRP) channel by tyrosine phosphorylation. SRC family kinase-dependent tyrosine phosphorylation of TRPV4 on TYR-253 mediates its response to hypotonic stress. J Biol Chem. 2003;278:11520–7. doi: 10.1074/jbc.M211061200. [DOI] [PubMed] [Google Scholar]
- 90.Sternfeld L, Krause E, Schmid A, Anderie I, Latas A, Al-Shaldi H, Kohl A, Evers K, Hofer HW, Schulz I. Tyrosine phosphatase PTP1B interacts with TRPV6 in vivo and plays a role in TRPV6-mediated calcium influx in HEK293 cells. Cell Signal. 2005;17:951–60. doi: 10.1016/j.cellsig.2004.11.012. [DOI] [PubMed] [Google Scholar]
- 91.Vazquez G, Wedel BJ, Kawasaki BT, Bird GS, Putney JW., Jr Obligatory role of Src kinase in the signaling mechanism for TRPC3 cation channels. J Biol Chem. 2004;279:40521–8. doi: 10.1074/jbc.M405280200. [DOI] [PubMed] [Google Scholar]
- 92.Kawasaki BT, Liao Y, Birnbaumer L. Role of Src in C3 transient receptor potential channel function and evidence for a heterogeneous makeup of receptor- and store-operated Ca2+ entry channels. Proc Natl Acad Sci U S A. 2006;103:335–40. doi: 10.1073/pnas.0508030102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Hisatsune C, Kuroda Y, Nakamura K, Inoue T, Nakamura T, Michikawa T, Mizutani A, Mikoshiba K. Regulation of TRPC6 channel activity by tyrosine phosphorylation. J Biol Chem. 2004;279:18887–94. doi: 10.1074/jbc.M311274200. [DOI] [PubMed] [Google Scholar]
- 94.Zhang W, Tong Q, Conrad K, Wozney J, Cheung JY, Miller BA. Regulation of TRP channel TRPM2 by the tyrosine phosphatase PTPL1. Am J Physiol Cell Physiol. 2007;292:C1746–58. doi: 10.1152/ajpcell.00569.2006. [DOI] [PubMed] [Google Scholar]
- 95.Woo DH, Jung SJ, Zhu MH, Park CK, Kim YH, Oh SB, Lee CJ. Direct activation of transient receptor potential vanilloid 1(TRPV1) by diacylglycerol (DAG) Mol Pain. 2008;4:42. doi: 10.1186/1744-8069-4-42. 8069-4-42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Zhang X, Huang J, McNaughton PA. NGF rapidly increases membrane expression of TRPV1 heat-gated ion channels. EMBO J. 2005;24:4211–23. doi: 10.1038/sj.emboj.7600893. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Jeske NA, Patwardhan AM, Henry MA, Milam SB. Fibronectin stimulates TRPV1 translocation in primary sensory neurons. J Neurochem. 2009;108:591–600. doi: 10.1111/j.1471-4159.2008.05779.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Van Buren JJ, Bhat S, Rotello R, Pauza ME, Premkumar LS. Sensitization and translocation of TRPV1 by insulin and IGF-I. Mol Pain. 2005;1:17. doi: 10.1186/1744-8069-1-17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Staruschenko A, Jeske NA, Akopian AN. Contribution of TRPV1-TRPA1 interaction to the single channel properties of the TRPA1 channel. J Biol Chem. 2010;285:15167–7. doi: 10.1074/jbc.M110.106153. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Sadofsky LR, Sreekrishna KT, Lin Y, Schinaman R, Gorka K, Mantri Y, Haught JC, Huggins TG, Isfort RJ, Bascom CC, Morice AH. Unique Responses are Observed in Transient Receptor Potential Ankyrin 1 and Vanilloid 1 (TRPA1 and TRPV1) Co-Expressing Cells. Cells. 2014;3:616–2. doi: 10.3390/cells3020616. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Fischer MJ, Balasuriya D, Jeggle P, Goetze TA, McNaughton PA, Reeh PW, Edwardson JM. Direct evidence for functional TRPV1/TRPA1 heteromers. Pflugers Arch. 2014;466:2229–41. doi: 10.1007/s00424-014-1497-z. [DOI] [PubMed] [Google Scholar]
- 102.Gallo EM, Cante-Barrett K, Crabtree GR. Lymphocyte calcium signaling from membrane to nucleus. Nat Immunol. 2006;7:25–32. doi: 10.1038/ni1295. [DOI] [PubMed] [Google Scholar]
- 103.Christo SN, Diener KR, Hayball JD. The functional contribution of calcium ion flux heterogeneity in T cells. Immunol Cell Biol. 2015;93:694–70. doi: 10.1038/icb.2015.34. [DOI] [PubMed] [Google Scholar]
- 104.Smith JB, Knowlton RP, Agarwal SS. Human lymphocyte responses are enhanced by culture at 40 degrees C. J Immunol. 1978;121:691–4. [PubMed] [Google Scholar]
- 105.Hanson DF. Fever and the immune response. The effects of physiological temperatures on primary murine splenic T-cell responses in vitro. J Immunol. 1993;151:436–48. [PubMed] [Google Scholar]
- 106.Tournier JN, Hellmann AQ, Lesca G, Jouan A, Drouet E, Mathieu J. Fever-like thermal conditions regulate the activation of maturing dendritic cells. J Leukoc Biol. 2003;73:493–501. doi: 10.1189/jlb.1002506. [DOI] [PubMed] [Google Scholar]
- 107.Mace TA, Zhong L, Kilpatrick C, Zynda E, Lee CT, Capitano M, Minderman H, Repasky EA. Differentiation of CD8+ T cells into effector cells is enhanced by physiological range hyperthermia. J Leukoc Biol. 2011;90:951–62. doi: 10.1189/jlb.0511229. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Xiao B, Coste B, Mathur J, Patapoutian A. Temperature-dependent STIM1 activation induces Ca(2)+ influx and modulates gene expression. Nat Chem Biol. 2011;7:351–8. doi: 10.1038/nchembio.558. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Orrenius S, Zhivotovsky B, Nicotera P. Regulation of cell death: the calcium-apoptosis link. Nat Rev Mol Cell Biol. 2003;4:552–65. doi: 10.1038/nrm1150. [DOI] [PubMed] [Google Scholar]
- 110.Nilius B, Owsianik G. Transient receptor potential channelopathies. Pflugers Arch. 2010;460:437–50. doi: 10.1007/s00424-010-0788-2. [DOI] [PubMed] [Google Scholar]
- 111.Nilius B, Owsianik G, Voets T, Peters JA. Transient receptor potential cation channels in disease. Physiol Rev. 2007;87:165–217. doi: 10.1152/physrev.00021.2006. [DOI] [PubMed] [Google Scholar]
- 112.Sel S, Rost BR, Yildirim AO, Sel B, Kalwa H, Fehrenbach H, Renz H, Gudermann T, Dietrich A. Loss of classical transient receptor potential 6 channel reduces allergic airway response. Clin Exp Allergy. 2008;38:1548–5. doi: 10.1111/j.1365-2222.2008.03043.x. [DOI] [PubMed] [Google Scholar]
- 113.Vennekens R, Olausson J, Meissner M, Bloch W, Mathar I, Philipp SE, Schmitz F, Weissgerber P, Nilius B, Flockerzi V, Freichel M. Increased IgE-dependent mast cell activation and anaphylactic responses in mice lacking the calcium-activated nonselective cation channel TRPM4. Nat Immunol. 2007;8:312–20. doi: 10.1038/ni1441. [DOI] [PubMed] [Google Scholar]
- 114.Grace MS, Baxter M, Dubuis E, Birrell MA, Belvisi MG. Transient receptor potential (TRP) channels in the airway: role in airway disease. Br J Pharmacol. 2014;171:2593–607. doi: 10.1111/bph.12538. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Holzer P. Transient receptor potential (TRP) channels as drug targets for diseases of the digestive system. Pharmacol Ther. 2011;131:142–70. doi: 10.1016/j.pharmthera.2011.03.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Zielinska M, Jarmuz A, Wasilewski A, Salaga M, Fichna J. Role of transient receptor potential channels in intestinal inflammation and visceral pain: novel targets in inflammatory bowel diseases. Inflamm Bowel Dis. 2015;21:419–27. doi: 10.1097/MIB.0000000000000234. [DOI] [PubMed] [Google Scholar]
- 117.Kochukov MY, McNearney TA, Fu Y, Westlund KN. Thermosensitive TRP ion channels mediate cytosolic calcium response in human synoviocytes. Am J Physiol Cell Physiol. 2006;291:C424–32. doi: 10.1152/ajpcell.00553.2005. [DOI] [PubMed] [Google Scholar]
- 118.Xu SZ, Sukumar P, Zeng F, Li J, Jairaman A, English A, Naylor J, Ciurtin C, Majeed Y, Milligan CJ, Bahnasi YM, Al-Shawaf E, Porter KE, Jiang LH, Emery P, Sivaprasadarao A, Beech DJ. TRPC channel activation by extracellular thioredoxin. Nature. 2008;451:69–72. doi: 10.1038/nature06414. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Schattling B, Steinbach K, Thies E, Kruse M, Menigoz A, Ufer F, Flockerzi V, Bruck W, Pongs O, Vennekens R, Kneussel M, Freichel M, Merkler D, Friese MA. TRPM4 cation channel mediates axonal and neuronal degeneration in experimental autoimmune encephalomyelitis and multiple sclerosis. Nat Med. 2012;18:1805–11. doi: 10.1038/nm.3015. [DOI] [PubMed] [Google Scholar]
- 120.Sulk M, Seeliger S, Aubert J, Schwab VD, Cevikbas F, Rivier M, Nowak P, Voegel JJ, Buddenkotte J, Steinhoff M. Distribution and expression of non-neuronal transient receptor potential (TRPV) ion channels in rosacea. J Invest Dermatol. 2012;132:1253–62. doi: 10.1038/jid.2011.424. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Issa F, Schiopu A, Wood KJ. Role of T cells in graft rejection and transplantation tolerance. Expert Rev Clin Immunol. 2010;6:155–69. doi: 10.1586/eci.09.64. [DOI] [PubMed] [Google Scholar]
